ABSTRACT
Let G be a group such that is finite and simple. The non-commuting, non-generating graph of G has vertex set , with edges corresponding to pairs of elements that do not commute and do not generate G. Complementing our previous investigation of this graph for non-simple groups, we show that is connected with diameter at most 5, with smaller upper bounds for certain families of groups. Using these bounds, we then prove that when G is simple, the diameter of the complement of the generating graph of G has a tight upper bound of 4, with the exception of at most one group with a graph of diameter 5.
1. INTRODUCTION
The generating graph of a group G has been investigated by many authors (for example, [5, 6, 13]) and encodes information about the 2-generation properties of G: its vertex set is and its edges are the generating pairs for G. The complement of this graph has also been studied (for example, in [38]), and for G non-abelian, its edge set necessarily includes all pairs of commuting elements. The remaining (and more interesting) non-generating pairs are described by the non-commuting, non-generating graph of G, with vertex set and an edge if and only if and .
Since its introduction in [10], the study of the graph has been motivated by the importance of the 2-generation properties of groups throughout (abstract and computational) group theory and its applications, and by Cameron’s [9, Section 2.6] hierarchy of graphs. This is a sequence of graphs defined on G (for convenience, we will take the vertex set of each to be ) such that, as long as G is non-abelian, each is a spanning subgraph of the next graph in the sequence. The final three of these graphs are the commuting graph (whose edges are the commuting pairs of elements of G), the non-generating graph (that is, the complement of the generating graph) and the complete graph. The well-studied generating graph is the difference between the final two graphs. It is therefore natural to investigate the next difference, between the non-generating graph and the commuting graph, with all vertices corresponding to central elements deleted (these would otherwise necessarily be isolated). This difference is precisely .
It was proved in [10] that if G is nilpotent (or more generally, if all maximal subgroups of G are normal) and has an edge, then either the subgraph induced by the non-isolated vertices of is connected with diameter 2 or itself is connected with diameter 3. In our companion paper [22], we extend these results to the general case where is not simple. The only additional possibilities here are that G is infinite and has diameter 3 or 4, or that is the union of two connected components of diameter 2 (we provide in that paper a detailed structural description of the finite groups G satisfying this case).
In this paper, we address the case where is finite and (non-abelian) simple. Our main theorem assumes that G itself is simple. Here, denotes the diameter of , which is always at least 2, since no vertex is adjacent to any of its powers. Compared to Theorem 6.1 from the thesis [21] on which this paper is based, our theorem takes into account new information on the maximal subgroups of the monster group ; see Remark 2.7.
Theorem 1.1Let G be a non-abelian finite simple group.
is connected with diameter at most 5.
If , then . If instead G is any other sporadic simple group or the Tits group, then .
For certain simple groups G, Table 1 lists exact values or upper bounds for .
Table 1.Exact values or upper bounds for , for certain simple groups G
G
. | Conditions
. |
. |
---|
| | 2 |
| q even or | 2 |
| | 3 |
| q odd and | 3 |
| | 4 |
An | n even | |
An | n odd | |
| | |
| q odd | |
G
. | Conditions
. |
. |
---|
| | 2 |
| q even or | 2 |
| | 3 |
| q odd and | 3 |
| | 4 |
An | n even | |
An | n odd | |
| | |
| q odd | |
Table 1.Exact values or upper bounds for , for certain simple groups G
G
. | Conditions
. |
. |
---|
| | 2 |
| q even or | 2 |
| | 3 |
| q odd and | 3 |
| | 4 |
An | n even | |
An | n odd | |
| | |
| q odd | |
G
. | Conditions
. |
. |
---|
| | 2 |
| q even or | 2 |
| | 3 |
| q odd and | 3 |
| | 4 |
An | n even | |
An | n odd | |
| | |
| q odd | |
Remark 1.2Table 1 shows that there exist non-abelian finite simple groups G with equal to 2, 3 or 4. We note that the baby monster group , the monster group and are the only known simple groups G for which . Indeed, we do not know if infinitely many such groups exist, nor if is possible.
At the end of each of Sections 3.1, 3.3, 5.1 and 5.2, we provide further examples of diameters achieved by various simple groups, computed using Magma [3] (see Section 2.1 for a discussion of our computational methods). Notably, none of our examples meet the upper bounds in rows 6–9 of Table 1.
Continuing the comparison between and the generating graph of G from [10, 22], we note that in the non-abelian finite simple case, the diameter of the former graph may be larger than that of the latter, which is equal to 2 by [5, Theorem 1.2].
Remark 1.3It is easy to determine the connectedness and diameter of the non-commuting, non-generating graphs of certain infinite simple groups G. For instance, if G is a Tarski monster, then each nontrivial proper subgroup of G has a fixed prime order, and so each vertex of is isolated. Additionally, it is a straightforward consequence of [1, Proposition 2.1] that whenever G is not 2-generated. However, it is an open problem to investigate the connectedness and diameter of when the infinite simple group G is 2-generated and has an edge.
Remark 1.4Our proof of Theorem 1.1 involves Proposition 3.10, which corrects [20, Theorem 1.1] on the intersection graph (defined in Section 2.1), in the case of the monster group. Additionally, our proof uses Theorem 4.5, on subspace stabilizers in simple unitary groups. In [21, Ch. 3], we prove more general versions of that theorem to obtain lower bounds for the base sizes of primitive subspace actions of almost simple classical groups.
We also note that our proof of Theorem 1.1(i) relies on the classification of finite simple groups. In Section 2.2, we prove the following special case that does not depend on the classification.
Proposition 1.5Let G be a non-abelian finite simple group, and suppose that every maximal subgroup of G has even order. Then is connected with diameter at most 5.
In fact, it follows from the classification (see Theorem 2.6 and Remark 2.7) that in addition to a few sporadic groups, the finite simple groups containing maximal subgroups of odd order belong to only three infinite families. Thus the above proposition serves as a useful reduction for the proof of Theorem 1.1(i).
Now, if is simple, then is a vertex of whenever . Thus [22, Lemma 2.9(ii)] yields the following corollary of Theorem 1.1(i).
Corollary 1.6Let G be a group such that is finite and simple. Then is connected, with .
If instead is finite and almost simple but not simple, then we observe from [22, Theorems 1.2–1.3 and Lemma 6.5] that is connected with diameter 2 or 3. Magma computations show that and , and so both possibilities occur.
Finally, as a consequence of Theorem 1.1, we will prove the following analogue of that theorem for the non-generating graph of a non-abelian finite simple group G. Recall that this graph is the complement of the generating graph of G. Hence its vertex set is , and vertices x and y are adjacent if and only if . Note that is a spanning subgraph of (since ). As above, we write to denote the baby monster group and to denote the monster group. Similar to Theorem 1.1, we incorporate new information on the maximal subgroups of (compared to [21, Theorem 6.5.4]) and its intersection graph.
Theorem 1.7Let G be a non-abelian finite simple group. The non-generating graph is connected with diameter 4 or 5 if , at most 4 if and at most 3 if every maximal subgroup of G has even order. Furthermore, if , then .
An earlier version of this theorem was communicated by this paper’s author to Peter Cameron, and appears as a remark after Proposition 12.3 in his paper [9]. That version of the theorem is a key component of the proof of [38, Theorem 1] (see the proof of [38, Proposition 10]), which investigates the connectedness and diameter of for a finite group G. This graph is closely related to the soluble graph of G, recently studied in [8]; see Remark 7 in that paper.
The remainder of this paper is structured as follows. The focus of Section 2 is on preliminary results and includes a comment on our computational methods and a proof of Proposition 1.5. In Section 3, we prove the alternating, sporadic and exceptional cases of Theorem 1.1. We consider matrices and stabilizers of subspaces in linear and unitary groups in Section 4 and then apply our results in Section 5 to prove Theorem 1.1 for those groups. Additionally, Section 5 concludes with a proof of Theorem 1.7.
2. PRELIMINARIES
Here, we present preliminary results on the non-commuting, non-generating graph of a group and on maximal subgroups of finite simple groups; in particular, we prove Proposition 1.5.
2.1. The non-commuting, non-generating graph
In this subsection, G is an arbitrary group, unless specified otherwise. Given vertices x and y of a graph Γ, we denote the distance in Γ between x and y by , and if x and y are adjacent, then we write .
Lemma 2.1Let H be a proper non-abelian subgroup of G. Then the induced subgraph of corresponding to is connected with diameter 2.
Proposition 2.2Suppose that G is finite and insoluble. Then has no isolated vertices. Equivalently, each element of is a non-central element of some maximal subgroup.
Proof.It is clear from the definition of that a vertex is isolated if and only if x lies in a unique maximal subgroup M of G and . Since G is finite, it follows from [22, Proposition 2.6] that each isolated vertex of lies in an abelian maximal subgroup of G. However, each maximal subgroup of the finite insoluble group G is non-abelian [29].
Our next result highlights a connection between and the intersection graph of G. This latter graph, introduced in [14], has vertex set the proper nontrivial subgroups of G, with subgroups adjacent if and only if they intersect nontrivially.
Lemma 2.3Let Γ be equal to or the non-generating graph of G. Additionally, suppose that and that Γ is connected with a finite diameter. Then .
Proof.By Propositions 12.1 and 2.3 of [9] (which assume that G is finite, but whose proofs also hold for infinite groups), the diameters of and the non-generating graph of G differ by at most one. As Z(G) is trivial, is a spanning subgraph of the non-generating graph of G, and so the diameter of the former is at least the diameter of the latter.
We conclude this subsection with a brief description of our computational methods for determining in Magma (for sufficiently small G). Let X be a set of representatives of generators for the non-central cyclic subgroups of G, and for each , let Yx be a set of representatives for the elements of X up to conjugacy in . We also obtain a set Xʹ from X by identifying elements if they generate conjugate cyclic subgroups. It is easy to see that contains and that , where Γ is the graph obtained from by identifying vertices that generate identical cyclic subgroups. We obtain by calculating the maximum distance in Γ between the elements in each pair (x, y) with and , via computations involving centralizers of elements and intersections of maximal subgroups.
The Magma code used in these calculations is contained in the file diam_nc_ng in [19]; see also [21, Appendix A] for further information. When using this code to calculate in this paper, we always represent G as a permutation group of minimal degree. We note that the time and memory requirements for running this code are negligible for sufficiently small groups (such as A5), but are in the order of a few days and a few gigabytes for the largest groups investigated in this paper via these methods, namely, , and .
2.2. Maximal subgroups of non-abelian finite simple groups
Assume throughout this subsection that G is non-abelian, finite and simple.
Lemma 2.4Let L and M be maximal subgroups of G, with even.
If Z(L) contains an involution a, then contains a G-conjugate of a.
L contains an involution that does not lie in .
Suppose that is even, and let a be an involution of . Then contains an involution b that does not commute with a.
Suppose that lies in Z(M) and contains an involution f. Additionally, let . Then contains an involution fʹ with .
Proof.(i) Suppose that Z(L) contains an involution a, so that . By [25, Corollary 1], there exists such that . In fact, , as . Thus , and so the involution ag is non-central in L.
(ii) We will prove that L contains an involution . By (i), there exists an involution . If , then we can set s = y. If instead (so that ), then . Additionally, since Z(L) and are proper subgroups of L, there exists , and yh is a non-central involution of L. Moreover, . Therefore, , and so we can set .
(iii) Suppose first that is odd, let S be the set of involutions of M and let be the subgroup generated by S. If the involution commutes with each element of S, then , and so , contradicting the simplicity of G. Thus there exists . As , we can set b = r.
If instead Z(M) contains an involution z, then by applying (i) to M, we deduce that contains an involution c. If , then we can set b = c. Otherwise, as , we see that . Thus we may set b to be the involution .
(iv) Let s be the involution from (ii), so that and . If , then we can set . Assume therefore that . Since and similarly , we obtain , and hence . If , then contains , which contains f. This contradicts , and so we can set .
The following lemma is one of the main ingredients in the proof of Theorem 1.1.
Lemma 2.5Let L and M be maximal subgroups of G of even order, and let and . Then , as vertices of . Moreover, if L contains an involution a such that , then .
Proof.We will use Lemma 2.1 several times in this proof without further reference. First, we may assume that , as otherwise . Let be the set of involutions in . Lemma 2.4(ii) shows that there exists . Notice that . If , then , and so . If instead , then Lemma 2.4(iii) yields an involution such that . As is dihedral, . Since , we conclude in this case that , with if .
Now, let be the set of involutions with . By the previous paragraph, it remains to consider the case where and . Note that in this case.
Let , so that . Then , and since , it follows from the definition of that and . Thus . Lemma 2.4(iv) (with f = a and u = x) now shows that , as otherwise would contain an involution in . Hence contains an element not centralizing M and the element a not centralizing x. As no group is the union of two proper subgroups, there exists centralizing neither M nor x. Thus and , and therefore .
For each , Proposition 2.2 gives for some maximal subgroup K of G (of even order). Thus by Lemma 2.5.
As the proofs of [25, Corollary 1] and Proposition 2.2 are independent of the classification of finite simple groups, so are the proofs of Lemmas 2.4 and 2.5 and Proposition 1.5.
To prove Theorem 1.1(i), it remains to consider the non-abelian finite simple groups that contain maximal subgroups of odd order.
Theorem 2.6Suppose that G is not isomorphic to the monster group. Then G has a maximal subgroup of odd order if and only if G is isomorphic to one of the following:
Ap, with p being a prime, and ;
, with n being a prime, and if n = 2;
, with n an odd prime and ; or
the Mathieu group , the Thompson group or the baby monster group .
Remark 2.7Let G be the monster group. By [35, Theorem 2], any maximal subgroup of G of odd order has shape 59:29 or 71:35. It was claimed in [31, 32] that maximal subgroups of G isomorphic to and were constructed computationally (using data that were not made publicly available). This suggested that no maximal subgroup of shape 59:29 or 71:35 exists. However, a brand new computational paper [17] shows (via publicly available code) that, up to conjugacy, G has unique maximal subgroups of shape 59:29 and , but no subgroup isomorphic to . Hence, up to conjugacy, 59:29 is the unique maximal subgroup of G of odd order.
3. ALTERNATING, SPORADIC AND EXCEPTIONAL SIMPLE GROUPS
In this section, we prove the alternating, sporadic and exceptional cases of Theorem 1.1. Note that we consider the Tits group together with the sporadic groups.
3.1. Alternating groups
Let G be the alternating group An of degree . We will prove the following result by considering the standard action of G on the set .
Theorem 3.1The graph of the alternating group G of degree is connected with diameter at most 4 or at most 3 if n is even.
Recall that a derangement of G is an element with no fixed points.
Lemma 3.2Let x be a derangement of G such that acts intransitively on Ω. In addition, let and be 2-subsets of distinct orbits of x on Ω. Finally, if x has exactly two orbits on Ω, then let , and otherwise, let . Then .
Proof.If has exactly two orbits on Ω, then at least one of these orbits has length at least three. Hence, in each case, , and so . Additionally, acts intransitively on Ω and is therefore a proper subgroup of G. Thus .
Lemma 3.3Let . Then . If , then at least one of x and y is a derangement, and if , then both are derangements.
Proof.Let α and β be distinct points in Ω. The point stabilizer is a maximal subgroup of G. As , it follows from Lemma 2.1 that for all . Additionally, is maximal in each of Gα and Gβ. Let and , such that . Then , as otherwise would centralize g. Similarly, . Hence some element of H centralizes neither g nor k, and it follows that .
To complete the proof, we will show that each derangement is adjacent in to some non-derangement. This follows immediately from Lemma 3.2 if is intransitive on Ω. Assume therefore that x is an n-cycle , with for each i, and n is odd. It suffices to prove that there exists with ; it will then follow that and , so that . For each n-cycle with , there exists an element such that . There are of these n-cycles, where ϕ is Euler’s totient function. We deduce that there exist such that and for some . Since , we can choose .
We will write to denote the support of an element .
Lemma 3.4Let x and y be derangements of G such that each of and acts intransitively on Ω. Then .
Proof.Let and be two subsets of distinct orbits of on Ω, such that , and let and be 2-subsets of distinct orbits of on Ω. We may assume that . If n = 5, then each of and has exactly two orbits on Ω.
Case (a): and each have exactly two orbits on Ω. We see from Lemma 3.2 that and . In particular, if a = b, then .
Assume therefore that , and hence either or . If , then either and ; or . As , we can repeat the argument with α2 replaced by an element of to obtain . Since is dihedral, is a path in and .
Case (b): and each have at least three orbits on Ω. By Lemma 3.2, and . Observe that , and so . Hence if , then is a path in , and so . If instead , then t = 3 and , hence and .
Case (c): Exactly one of and has exactly two orbits on Ω. We may assume that has exactly two orbits on Ω. Lemma 3.2 shows that and . Observe that and is intransitive. Hence and .
Lemma 3.3 shows that is connected with diameter at most 4. Furthermore, if n is even, then each derangement of G generates a subgroup that acts intransitively on Ω. Hence Lemmas 3.3 and 3.4 yield .
Magma calculations show that whenever . We do not know if there exists n > 10 such that .
3.2. Sporadic groups
Let G be a sporadic finite simple group, or the Tits group . In this subsection, we implicitly use [12, 43] to determine information about maximal subgroups, conjugacy classes and centralizers in G. In particular, all maximal subgroups of the sporadic groups are listed in [43] (up to conjugacy), with the exceptions of the maximal subgroups of the monster group isomorphic to and constructed in [18] and the maximal subgroups isomorphic to 59:29 constructed in [17] (see the introductions of those last two papers for further details). Additionally, the subgroup listed in [43] is not in fact a subgroup of ; see Remark 2.7.
Theorem 3.5The graph of the sporadic group or Tits group G is connected with diameter 4 or 5 if and at most 4 otherwise. If , then is given in Table 1.
A few of the proofs in this subsection, where specified, involve straightforward computations in GAP [24] that utilize the Character Table Library [4], with runtimes of at most a few seconds. This library contains the necessary information regarding all maximal subgroups of G, except for certain maximal subgroups of the monster group.
Proposition 3.6Each maximal subgroup of G has a centre of order at most 2.
Proof.For each maximal subgroup M of not included in the GAP Character Table Library, no element of has a centralizer of order , hence . For each other maximal subgroup of each G, we verify the result using GAP.
In light of Lemma 2.5, it will be useful to determine neighbours of involutions in .
Proposition 3.7Let x be an element of G of order at least 3, and suppose that x lies in a maximal subgroup of G of even order. Then x is adjacent in to some involution of G.
Proof.It suffices to show that there exists a maximal subgroup of G that contains both x and an involution that does not centralize x. Let M be a maximal subgroup of G of even order, with . Since , Proposition 3.6 shows that . Let TM be the set of involutions of M. If , or if the normal subgroup of M is equal to M, then x does not centralize every involution of M. Note that this second condition holds whenever M is simple.
Using GAP, we see that if x does not lie in any simple maximal subgroup, nor in any maximal subgroup M with , then one of the following holds:
and ;
and ;
and ;
and ; or
and .
Our computations here involve the fusion of M-conjugacy classes in G. As such, when , we consider only the set of maximal subgroups of for which this fusion is stored in GAP (this is a subset of the set returned by the NamesOfFusionSources function). Similarly, when , our computations exclude the maximal subgroups of shape , for which this fusion is not stored.
Now, in case (iii), x is an element of the G-conjugacy class labelled in the atlas [12] and lies in a maximal subgroup , which clearly satisfies . In all other cases, is odd and , and hence x does not commute with any involution of M.
Lemma 3.8Suppose that , and let R and M be maximal subgroups of G isomorphic to . Additionally, let with . Then there exists such that .
Proof.Let . If , then we can set f = s. Otherwise, , and since , there exists . As , we observe that .
Lemma 3.9Suppose that , and let , with . Then .
Proof.The element s generates a Sylow 23-subgroup S of G and lies in maximal subgroups and . Each maximal subgroup of G that contains s and is not G-conjugate to R or T has shape 47:23. Moreover, and (using Proposition 3.6 and the atlas) . In addition, Proposition 2.2 shows that for some maximal subgroup L of G. We will show that whenever L satisfies certain properties and that we may often choose L to satisfy those properties.
Suppose first that and , so that . We may assume that , else by Lemma 2.1. If there exists a non-identity element that lies in then (if , say, then implies that , and then yields ). Note also that is a proper subset of , since . Thus regardless of whether such an element u exists, contains an element . Moreover, any such c satisfies by Lemma 2.1. In particular, if , then we can set c = s, and so . Assume therefore that . We claim that . Indeed, if , then c centralizes s and hence lies in . Thus for some , where z is the unique involution of Z(T). This implies that , and so , a contradiction. Therefore, and .
Next, suppose that , and . Then , while (since no L satisfying the given assumptions has order divisible by ). As , each is adjacent to s in . Moreover, by Lemma 2.1, and so again .
Assume now that . Using GAP, we see that there exists a maximal subgroup K of G containing y, such that either and ; or , and . Moreover, Proposition 3.6 and Lemma 2.4(i) show that contains a G-conjugate of y. Hence we can choose L to be a G-conjugate of K, and by the previous two paragraphs. To complete the proof, we will consider the remaining elements y.
Case (a): . Here, , and each maximal subgroup of G containing y has shape 47:23. As S is a Sylow subgroup of G, it follows that for some in a G-conjugate Sʹ of S. To complete the proof in this case, we will show that .
The element sʹ lies in a G-conjugate M of R. If , then it is clear that . Otherwise, , and so if s or sʹ lies in , then . Suppose finally that . We deduce from Lemma 3.8 that for some . Hence there exists centralizing neither s nor sʹ, and is a path in .
Case (b): . There is a unique G-conjugacy class of elements of order , and so we may choose if or if . In either case, no involution of G has a centralizer of order , and so Proposition 3.6 yields .
First suppose that there exists . Then , as does not divide and . Hence Lemma 2.1 yields . Additionally, , and so . Thus either or , and .
Suppose finally that . Each of G and R has a unique conjugacy class of elements of order 35, and L also contains elements of order 35. Hence there exists a G-conjugate M of R such that contains an element m of order 35, and since . Additionally, Lemma 3.8 shows that for some . Note that , since and . Thus and .
Recall that the vertices of the intersection graph of G are the proper nontrivial subgroups of G, and the graph’s edges are the pairs of subgroups that intersect nontrivially. The result [20, Theorem 1.1] on the diameter of (for all non-abelian finite simple groups G) does not take into account the new information about mentioned in Remark 2.7. We now correct that theorem in the case ; this will be useful in the proofs of Theorems 3.5 and 1.7.
Proposition 3.10Suppose that . Then the intersection graph of G is connected with diameter 5 or 6.
Proof.Let M1 and M2 be maximal subgroups of G, with being odd. By elementary arguments from the proof of [20, Theorem 1.1], it suffices to bound , where S1 and S2 are proper nontrivial subgroups of M1 and M2, respectively, such that S1 lies in no maximal subgroup of even order. By Remark 2.7, M1 has shape . Let S and H be subgroups of M1 of order 59 and 29, respectively, so that H is a Sylow subgroup of G.
We observe from [12] that . Arguing almost exactly as in the case of the proof of [20, Theorem 1.1], with the aid of computations involving the GAP Character Table Library, we determine that M1 is the unique maximal subgroup of G containing S; that the maximal subgroups of G containing H are precisely 84 conjugates of M1, three conjugates of , and one conjugate of ; and that for some . In particular, we may assume that .
Now, let R be a maximal subgroup of even order containing H. If S2 is not conjugate to S, then M2 can be chosen to have even order. Observe that the subgroup D generated by an involution of R and an involution of M2 is dihedral. Hence , and we obtain . If instead for some , then we similarly deduce that for some dihedral group , and thus . Therefore, .
Remark 3.11Let G, S and 59:29 be as in the proof of Proposition 3.10, and let F be a maximal subgroup of G of shape . It follows from the proof of Proposition 3.10 that if and only if there exists such that . Since M1 is the unique neighbour of S1 in , this occurs if and only if .
One possible obstruction to this occurring is the existence of elements such that . Since H lies in a unique conjugate of F, and M1 contains 59 conjugates of H, exactly 59 conjugates of F are adjacent to M1 in . An elementary counting argument now shows that exactly conjugates of M1 are adjacent to F. Additionally, by the proof of [37, Theorem 1.3], the number β of conjugates of F adjacent to F is at most , where R is a set of representatives for the G-conjugacy classes of elements of prime order. We calculate using the GAP Character Table Library that . We therefore deduce that there are at most conjugates of M1 such that for some . This is unfortunately not a useful upper bound, as it is larger than the number of conjugates of M1.
Determining a better estimate for β is difficult due to the large index of F in G. However, if a sufficiently small upper bound for β can be obtained, then the above argument (along with similar arguments for the other types of path between M1 and of length at most three) would yield the existence of an element x such that , so that . Otherwise, deducing the exact value of would require more detailed information about the neighbourhoods of conjugates of M1, F and the maximal subgroup of G. The large order of G is again a significant obstacle to obtaining this information.
For each , we obtain using Magma. In all other cases, for , let xi be an element of that lies in a maximal subgroup of even order, say Li. By Proposition 3.7, for some involution ai of G, and so (using Proposition 2.2 if ) we may choose Li so that . Lemma 2.5 now gives . Hence if each maximal subgroup of G has even order.
It remains to consider the case where G has a maximal subgroup K of odd order. By Theorem 2.6 and Remark 2.7, . Let such that h lies in no maximal subgroup of even order.
First assume that . Then 31:15. Let . If , then m lies in a maximal subgroup of G of shape , and otherwise is even. Thus each element of G lies in a maximal subgroup of even order, and so the element h does not exist. Hence the first paragraph of the proof yields .
Suppose next that . In this case, 47:23, and . Thus for each of order 23. As by Lemma 3.9, we conclude that . It follows that . In addition, the intersection graph of G has diameter 5 [20, Theorem 1.1], and so Lemma 2.3 yields . Therefore, .
Finally, assume that , so that 59:29, and . Thus for each of order 29. Let R be a maximal subgroup of G of even order containing s, and note that is a Sylow subgroup of G. If g is not conjugate to h, then, as in the first paragraph of the proof, for some involution a of G, there exists a maximal subgroup L of G such that . By Lemma 2.4(iii) and the fact that two involutions generate a dihedral group, R contains an involution b such that . Moreover, , which implies that . Hence , and so . If instead for some , then we similarly deduce the existence of involutions aʹ and bʹ such that , and so . Therefore, . As the intersection graph of G has diameter at least 5 by Proposition 3.10, Lemma 2.3 implies that .
Determining the exact diameter of (and similarly for the non-generating graph of in Theorem 1.7) is difficult for reasons similar to those discussed in Remark 3.11.
3.3. Exceptional groups of Lie type
We now prove Theorem 1.1 in the case where G is an exceptional group of Lie type. In view of Proposition 1.5 and Theorem 2.6, it remains to prove the following. Here, q denotes a prime power, and we do not include ; for that group, see Theorem 5.1.
Theorem 3.12Let G be a finite simple group. If , or if q is odd and , then is connected with diameter at most 4.
Given a group X, let QX be the (characteristic) subgroup generated by the involutions of X.
Lemma 3.13Let , and let H be a quasisimple group containing a subgroup K, such that and . Additionally, let R be the central product , and let π be the natural epimorphism . If , then q = 3, QR is a maximal subgroup of R that contains , and .
Proof.We observe that and are normal subgroups of and and are normal subgroups of R, with , and . Additionally, and . Furthermore, the centres of R, J, , and all coincide, and this common centre of order is equal to . We divide the rest of the proof into two cases.
Case (a): . We first show that . This is clear if q = 2, so assume otherwise. Considering K as a copy of , let , where I2 is the identity matrix. The choices for A and B include the matrix
$$
for each
. As
S is quasisimple, and a quasisimple group’s proper normal subgroups are precisely its central subgroups, we see that
projects onto each of
S and
K and that the kernel of each projection map is isomorphic to
. Thus
. Since the image under
π of each element of
L is an involution, it follows that
.
Now, since is quasisimple and , no proper normal subgroup of contains . Hence no proper normal subgroup of R contains J. As , we obtain .
Case (b): q = 3. Here, QJ is the image under π of the direct product of the normal Sylow 2-subgroups of S and K, which are isomorphic to the quaternion group Q8. Thus is maximal in and . As the normal subgroup of contains , it follows that . Since , we deduce that .
If , then . We will therefore assume that . Then is the maximal subgroup of , and , which is maximal in . Since , we obtain The centralizer is equal to , and hence so is .
For an example where , take H to be the quasisimple group .
Proposition 3.14Let H be a finite quasisimple group. If , then assume that is not isomorphic to A7 or to for any odd q. Then .
Proof.This is a straightforward consequence of the main theorem of [27], which implies that there exists an involution if is even. Clearly, a also exists if is odd. Each proper normal subgroup of the quasisimple group H lies in Z(H), and so .
Lemma 3.15Let G be a simple group in with q odd (so that if ), let a be an involution of G and let . Then .
Proof.We observe from [26, pp. 171–177] that , and
We will divide the remainder of the proof into the above cases. In each case, since , it remains to show that . Note that this is certainly the case if .
Cases (a)–(c). By applying Proposition 3.14 to Y in Case (b) and J in Case (c), we easily deduce that in each case.
Case (d). If , then [42, p. 125], and . Suppose therefore that and that . Notice that QY contains , and thus .
In order to apply Lemma 3.13, we will show that the quasisimple group H contains a subgroup K such that and . Such K clearly exists if there exists with and for some positive integer i. Hence this is the case if . If instead , then H contains a subgroup , with [28, p. 927, item (3)]. Finally, if , then let V1, V2 and V3 be non-degenerate two-dimensional subspaces of the symplectic space , such that . Then the intersection of the stabilizers in H of V1, V2 and V3 is isomorphic to . The diagonal subgroup of this direct product is isomorphic to and has centre Z(H).
Lemma 3.13 now shows that QR is a maximal subgroup of R, that and that . It follows from this first fact that the proper subgroup QY of Y is maximal and equal to . Since and , each inner automorphism of QR acts on as an inner automorphism of . Thus the outer automorphism of induced by A does not extend to an inner automorphism of QR. Therefore, , which is equal to Z(Y) from the start of the proof. As , we conclude that .
Theorem 2.6 implies that each maximal subgroup of G has even order, and so it suffices by Proposition 2.2 and Lemma 2.5 to show that each non-involution is adjacent in to some involution. Let M be a maximal subgroup of G containing x. Assume first that , so that q is odd, and let a be an involution of M. If , then we are done. Otherwise, . By Lemma 3.15, a is the unique non-identity element of that centralizes each involution of . Hence for some involution .
Assume now that . By Theorem 4 on page 122 of [41], is nilpotent for each . As the only non-abelian finite simple groups with nilpotent maximal subgroups are for various primes p (see [40, p. 183]), we deduce that and . If , for some proper power q0 of 2 dividing q, then . It follows in this case that M contains an involution a not centralizing x, and so . If instead , then by Section 4, page 133 and Theorem 9 on page 137 of [41], M is conjugate in G to
If , then again and for some involution . If instead , then let N be the normal subgroup of M. As M is Frobenius, there exists a complement H of N with . No non-identity element of H centralizes x, and so x is adjacent in to the unique involution of H.
Suppose finally that . If , then x centralizes no non-identity element of S, and so x is adjacent in to each involution of S. Now, Z(S) consists of all elements of S of order at most 2, and each element of has order 4, by Lemma 1 on page 112 and Theorem 7 on page 133 of [41]. Additionally, G has a unique conjugacy class of cyclic subgroups of order 4 (see the final sentence of page 121 and Proposition 18 on page 126 of [41]), and so if the non-involution x lies in S, then it also lies in a maximal subgroup . By the previous paragraph, x is adjacent in to an involution of L.
Magma calculations show that and . We do not know if any finite simple exceptional group G satisfies . Observe also that the proof of Theorem 3.12 relies on useful properties of the involution centralizers in G (and of and its maximal subgroups). Indeed, in the q odd case, we have considered precisely those finite simple groups whose involutions are all centralized by maximal subgroups of maximal rank (see [26, Theorem 4.5.1] and [36]). The involution centralizers for the remaining exceptional groups have more complicated structures (see for example [2] and [26, Theorem 4.5.1]), so alternative methods would be necessary to prove an analogue of Theorem 3.12 for those groups.
4. Matrices and subspace stabilizers
We now consider various matrices and subspace stabilizers. The results here will be useful when proving Theorem 1.1 for linear and unitary groups. Let n be a positive integer, q a prime power and Ik the k × k identity matrix over , for a positive integer k. Additionally, let be the n × n matrix with (i, j) entry equal to 1 and all other entries equal to 0.
4.1. Matrices and subspace stabilizers in linear groups
Definition 4.1Let
be a (monic) polynomial in
. The
companion matrix of
f is the
n ×
n matrix
where
$A:=()$
,
$B:=()$
and
when
n = 1. If
f is irreducible over
, then for each positive integer
k, the
hypercompanion matrix of
f k is the
kn ×
kn matrix
Since a subgroup of is irreducible if and only if the characteristic polynomial χA of A is irreducible, the following lemma will shed light on certain irreducible subgroups.
Lemma 4.2Let . Then the set of irreducible factors in of the binomial is .
Proof.For each positive integer k, let ϕk be the homomorphism of , and let Hk be the image of ϕk, of order . Additionally, let and . Then . Since , there are exactly t distinct roots of the binomial . Hence , and
It remains to show that the binomial is irreducible over for each i. Since , we see that divides . Additionally, if r is divisible by 4, then so is q − 1. It therefore suffices by [34, Theorems 3.3 and 3.35] to prove that r is coprime to .
Since , there exists with , and hence . Thus , and so r divides . Hence .
Recall that matrices are similar if and only if .
Lemma 4.3Let , and suppose that and that acts irreducibly on . Then . If, in addition, n = 2 and , then .
Proof.Since , there exists such that . Thus χA, which is irreducible over , divides . Furthermore, is the splitting field for χA over , and A is similar to the companion matrix . To prove the required result, we will study χA as a polynomial in . By Lemma 4.2, the set of irreducible factors in of is .
If n is odd, then χA is irreducible over . Thus , and for some , with . Since , we deduce that b = 1. Hence a = 1 and n = 1.
If instead
n is even, then each irreducible factor of
χA in
has degree
. In particular,
, and there exist
such that
with
and
. Here,
, and thus
. Therefore,
, and hence
. In particular, if
q is even, then
has odd order, and
n = 2. Otherwise, arguing as in [
11, pp. 569–570] shows that the irreducible polynomial
cannot have degree 4, and again
n = 2.
Finally, if n = 2 and , then for some . Hence the irreducible polynomial χA divides . It follows from Lemma 4.2 and the previous paragraph that , which is irreducible if and only if .
Let , with and if n = 2, so that is simple. Additionally, let be the natural module for H and let HX be the stabilizer in H of a subspace X of V.
Lemma 4.4Let X and Y be distinct one-dimensional subspaces of V. Then:
Proof.Up to conjugacy by a fixed element of H, the stabilizer HX consists of the matrices in H with a first row of the form
$$
, with
, and
HY is the set of matrices with a second row of the form
$$
, with
. The restriction of a matrix in
HX to its (1, 1) entry is a homomorphism from
HX to the abelian group
, and its kernel clearly contains all commutators in
HX. Thus (i) holds. We split the proof of (ii) and (iii) into two cases. In each case, let
.
Case (a): n = 2. Let ω be a primitive element of , and . The equality AW = WA gives , which is abelian. Hence we obtain (iii). If , then DW = WD, and we deduce that , hence (ii).
Case (b): . It is easy to check the result if n = 3 and q = 2, so assume otherwise. If n = 3, then let and . We deduce from the equalities BW = WB and that W is diagonal. If instead , then let r and s be distinct integers with and . Then , and the equalities (for all r and s) imply that W is again diagonal.
In each case, let be the matrix obtained from In by setting all entries in the final row to 1. As W is diagonal, the equality WF = FW yields . Thus we have proved (iii), and (ii) follows immediately.
4.2. Subspace stabilizers in unitary groups
Throughout this subsection, let , with if n = 3, so that is simple. Additionally, let V be the natural module for , equipped with the unitary form with Gram matrix In. Then a matrix lies in if and only if , where σ is the automorphism of . The following theorem will be useful when considering . Here, denotes the pointwise stabilizer in G of a set Δ of subspaces of V.
Theorem 4.5Let Δ be a set of k-dimensional subspaces of V.
An analogue of the following lemma holds for certain classical algebraic groups; see [7, Section 4.1].
Lemma 4.6Let W be an -dimensional subspace of V. Then contains a non-scalar matrix that acts as a scalar on W if and only if is totally singular or .
Proof.First, has two orbits on the set of one-dimensional subspaces of V: the non-degenerate subspaces and the totally singular subspaces (cf. Witt’s Lemma). Thus it suffices to consider two choices for W, one for each possible orbit containing . Let be the basis for V corresponding to the Gram matrix In, and let ω be a primitive element of .
Case (a): is non-degenerate. We may assume that , so that . Any matrix in that acts as a scalar on W also stabilizes and hence is equal to for some . Moreover, if and only if .
Let . Observe that if , then Bλ is a non-scalar matrix in that acts as λ on W. If instead , then any scalar satisfying also satisfies . Therefore, in this case, no non-scalar matrix in acts as a scalar on W.
Case (b): is totally singular. If q is even, then let and , and otherwise, let and . Then for all q, and so is singular. Hence we may assume that , so that . The direct sum of
$\left( \right)$
and
is a non-scalar matrix in
that acts trivially on
W.
Corollary 4.7Let U be an -dimensional subspace of V. Then contains a non-scalar matrix that acts as a scalar on U.
Proof.Observe that contains a one-dimensional totally singular subspace X. By Lemma 4.6, some non-scalar matrix in acts as a scalar on and hence on .
If , then the subspace of V lies in an -dimensional subspace W. Lemma 4.6 shows that if , then contains a non-scalar matrix that acts as a scalar on W and hence on . Therefore, .
Finally, assume that or that and . Then lies in an -dimensional subspace U of V. By Corollary 4.7, contains a non-scalar matrix that acts as a scalar on U and hence on . We again conclude that .
5. LINEAR AND UNITARY GROUPS AND THE NON-GENERATING GRAPH
In this section, we complete the proof of Theorem 1.1 by showing that it holds for linear and unitary groups. In addition, we prove Theorem 1.7. Define n, q, Ik and as in Section 4, and let , and . We will largely work in the group H, appealing to the following consequence of the definition of : for , the vertices and of are adjacent if and only if and .
5.1. Linear groups
Assume that , with and if n = 2, so that G is simple.
Theorem 5.1Let G be the linear group . Then . Moreover, if n = 2, then if q is even or , and otherwise.
Let be the set of one-dimensional subspaces of the natural module for H. We proceed by considering paths in containing elements that stabilize subspaces in .
Lemma 5.2Let , and . Then .
Proof.We may assume that . We claim that there exists a path in , where , , and . Since by Lemma 4.4(ii), the claim follows from Lemma 2.1 if K or L lies in . Otherwise, Lemma 4.4(iii) implies that and are proper subgroups of . Thus there exists an centralizing neither K nor L, and we can set .
Now, for all , and so by Lemma 4.4(i). Moreover, lies in HX or HY. Hence is a path in .
Lemma 5.3Suppose that n > 2 or , and let such that stabilizes no subspace in . Then there exist and such that .
Proof.We divide the proof into two cases.
Case (a): acts reducibly on V. It is clear that n > 2. The matrix A is similar to a block matrix whose final row of blocks is
$$
, with
R being a hypercompanion matrix and a proper submatrix of
A (see [
39, Theorem 8.10]). Since
, we may assume without loss of generality that
A is this block matrix.
Let . Then , where is spanned by
$ \in V$
. Each matrix in
can be viewed as a block matrix, with a final row
$$
of blocks, for some nonzero block
S with the same dimensions as
R. Hence
.
It remains to prove that . For a contradiction, suppose otherwise, so that for some . Then . As K has eigenvalue 1 with algebraic multiplicity n, while cK−1 has eigenvalue c, we deduce that . However, and , a contradiction.
Case (b): acts irreducibly on V. By [15, Proposition 3.6.1], A lies in a Singer subgroup S of , that is, a cyclic subgroup of order . Additionally, is the group of order , where is similar to the companion matrix and satisfies for each (see [33, pp. 187–188] and [30, p. 497]). As stabilizes the subspace in spanned by the all-ones vector, B also stabilizes some . Note that is equal to 1 if q is even or n is odd, and −1 otherwise. Additionally, N is soluble, and so its subgroup is proper in H. Furthermore, . Since , Lemma 4.3 and our assumptions on n and q yield .
It again suffices to find a matrix with and . If n is odd or q is even, then set K = B. If instead n = 2 and , then set for some square root of −1. Finally, if q is odd and n is even and greater than 2, then . Additionally, , which does not lie in by Lemma 4.3. Thus we set .
Let . We will bound . By Lemma 5.2, if and stabilize one-dimensional subspaces of V. Since , we may assume without loss of generality that stabilizes no such subspace. If , then Lemma 5.3 shows that for some with . Similarly, for some with . Hence , as required.
Suppose from now on that n = 2. Magma computations show that when q is odd and at most 9, so we will assume that q is even or at least 11. Let p be the prime dividing q, and let . In what follows, all information about elements and subgroups of G is from [16, Ch. XII], except where stated otherwise. We first note that and that each non-identity element of G lies in either a unique cyclic subgroup of order , a unique cyclic subgroup of order , or a unique Sylow p-subgroup isomorphic to the elementary abelian group E of order q. Moreover, all cyclic subgroups of G of a given order are conjugate. In addition, each element of order dividing q or lies in a parabolic maximal subgroup of shape , which is the stabilizer in G of a subspace in .
It now follows that divides . The unique cyclic subgroup of order containing x lies in a unique maximal subgroup . We divide the remainder of the proof into two cases.
Case (a): q is even. Suppose first that does not divide q − 1. All involutions of G are conjugate, as are all cyclic subgroups of order q + 1, and so y lies in a G-conjugate M of L. As , Lemma 2.1 yields if L = M. If instead , then contains an involution a [23, Lemma 2.3], and it is clear that . Hence and , and so .
Next, suppose that divides q − 1. Then y lies in a parabolic subgroup P of G of shape . As , there exists an involution . The centralizer of b in G is the Sylow 2-subgroup of P, and thus . Therefore, .
Case (b): q is odd. There exist G-conjugates J and K of L with [23, Lemma 2.2]. As above, J is the unique maximal subgroup of G containing an element s of order and K is the unique maximal subgroup containing a conjugate t of s. Neither s nor any neighbour of s in is a neighbour of t, and so . Hence it remains to show that .
Suppose first that divides . Notice that if , else . Additionally, since all cyclic subgroups of G of a given order are conjugate, so are all involutions of G. It follows that there exist G-conjugates R and S of L with and . If R = S, then by Lemma 2.1. Otherwise, as the dihedral group R is generated by a pair of non-commuting involutions, there exists an involution . In addition, Lemma 2.4(iii) yields an involution with . Since is dihedral, with and , we deduce that contains a path in from x to y. Hence .
Assume from now on that does not divide , so that y lies in a parabolic subgroup P of shape , with for some . If , then Lemma 5.3 yields for some with . Since by Lemma 5.2, we obtain .
Suppose therefore that , and let c be an element of P of order . Then is the pointwise stabilizer for some , and is dihedral of order q − 1 and equal to the setwise stabilizer . It is also clear that for each of the involutions , and in fact if , since . In addition, , and so if , then . Now, let cʹ be an element of of order , so that . Then . Since there are q cyclic subgroups of order in P, it follows that y has distance at most two from at least involutions of G. This in fact accounts for all involutions of G, since the centralizer in G of an involution is conjugate to L, which has order . Hence if , then . Otherwise, x is adjacent to each involution in , and so , completing the proof.
Using Magma, we see that has diameter 2 when and diameter 3 when . Note that the diameter of was incorrectly reported as 3 in [21].
5.2. Unitary groups
Assume that , with and if n = 3, so that G is simple. In fact, by Proposition 1.5 and Theorem 2.6, we may assume that n is prime.
Theorem 5.4Let G be the unitary group , with n being an odd prime.
Let V be the natural module for , and fix a basis for V so that the unitary form on V has Gram matrix In.
Lemma 5.5Suppose that n = 3, let ω be a primitive element of and let . Let , and . Additionally, let A1 and A2 be H-conjugates of B1.
Proof.Since and , we see that . Additionally, , and so , hence (i). For (ii), note that the matrices
$\left( \right)$
and
$\left( \right)$
lie in
H and conjugate
B1 to
B2 and
B3, respectively.
To prove (iii), we may assume that . Let such that , and note that B1 and A2 act as λ on the subspaces and EF of V, respectively. Hence stabilizes each subspace of , which has positive dimension since .
Finally, consists of diagonal matrices, and all H-conjugates of B1 have the same characteristic polynomial as B1. Since B1, B2 and B3 are the only diagonal matrices in with this characteristic polynomial, we obtain (iv).
Lemma 5.6Suppose that n is an odd prime.
The companion matrix is an element of H.
Let R be an element of H that is conjugate to . Then is adjacent in to an involution of G.
Let R and S be H-conjugates of the companion matrix . Then is at most 3 if n = 3 or at most 2 otherwise.
Proof.(i) This holds since and .
(ii) We may assume that . If q is odd, then let A be the direct sum of the matrices and , so that the (1, 1) and (n, n) entries of are equal to −1 and 1, respectively. If instead q is even, then let be the direct sum of the matrices and
$$
, so that the (1, 1) entry of
is equal to 0. In either case,
. Moreover,
R and
A are monomial, and so
. Thus
. As
A is an involution, so is
.
(iii) We may assume that and . Then R stabilizes the one-dimensional subspace X of V spanned by the all-ones vector. Let such that , and let , so that and . Additionally, let . It is easy to show that , and thus . Furthermore, as and L < H, it follows from Lemma 4.4(i) that if for some , then . Similarly, if for some , then . We split the remainder of the proof into three cases.
Case (a): , or n = 5 and . By Theorem 4.5, the pointwise stabilizer in H of contains a non-scalar matrix D. The previous paragraph shows that D centralizes neither R nor S, and in fact that .
Case (b): n = 5 and q = 4. Straightforward computations in Magma (with a runtime of about 1.5 h) show that is even (for all ), while (and hence ). It follows from the second last paragraph that for each involution .
Case (c): n = 3. For each , define as in Lemma 5.5. It is easy to check that , and since , R and Bi are monomial matrices, . Thus . Similarly, . Recall also from Lemma 5.5(ii)–(iii) that stabilizes a one-dimensional subspace of V. Hence, as in the third last paragraph, if for some i, then , and so . Otherwise, Lemma 5.5(iv) yields for some i, and so .
(i) Let . We will show that . By Proposition 2.2, each of x and y is non-central in some maximal subgroup. Using Lemma 2.5, we may assume that each maximal subgroup K of G with has odd order. By [35, Theorem 2], , where S is a Singer subgroup of H, that is, the intersection of H and a Singer subgroup of . It follows from [30, p. 497 and p. 512] that for some H-conjugate C of .
Now, by Lemma 5.6(ii), for some involution , and so for some maximal subgroup L of G of even order. As is prime, each non-identity element of lies in . Since , it follows that for some and some . Again using the fact that n is prime, it follows from Clifford’s Theorem that acts irreducibly on . Similar to Case (b) in the proof of Lemma 5.3 (but now working over ), we see that . By Lemma 4.3, , and so . Moreover, . Therefore, . It is now easy to see that for all i, and, in particular, .
If for some maximal subgroup M of G of even order, then since and , Lemma 2.5 shows that . Hence . Otherwise, y lies in a G-conjugate of . By the previous paragraph, for some H-conjugate D of C. Since and by Lemma 5.6(iii), we again obtain .
(ii) By [20, Remark 1.3] (see also [21, Ch. 4]), the intersection graph of G has diameter 5. Thus by Lemma 2.3. It remains to show that for all . In what follows, except where stated otherwise, all information about G is determined using Magma, via computations involving conjugacy classes of elements and subgroups of G, with a runtime of about 2 min.
If , then each of x and y is adjacent in to an involution, and Lemma 2.5 yields . Assume from now on that . Then x lies in no maximal subgroup of G of even order. As in the proof of (i), , where t is the image in G of some H-conjugate of . Since n > 3, it follows from Lemma 5.6(iii) that if , then .
Suppose finally that , and let . Then there exist maximal subgroups K and L of G such that , and either
In case (I), there exists an element that centralizes neither t nor K. Hence , and by Lemma 2.1, yielding . In case (II), we observe that and that for each non-identity . As by Lemma 2.1, we again obtain . We conclude in general that .
Using Magma, we observe that the non-commuting, non-generating graphs of , and have diameters 2, 2 and 3, respectively.
5.3. The non-generating graph of a non-abelian finite simple group
Let , and let M1 and M2 be maximal subgroups of G containing x and y, respectively. If and are even, then is dihedral for involutions and . Hence contains a path from x to y in , and . Thus we are done if every maximal subgroup of G has even order.
Assume now that is odd. Since is a spanning subgraph of , it follows from Theorems 2.6 and 1.1 that if G is not unitary of odd prime dimension, then is at most 5 if , and at most 4 otherwise.
Suppose therefore that , with n being an odd prime. As in the proof of Theorem 5.4, M1 is the image in G of the subgroup of , with S being a Singer subgroup of and C being a conjugate of the companion matrix . By Lemma 5.6(ii), there exists an involution such that the image c of C in G is adjacent to a in , and hence in .
Now, if M2 contains an involution b, then contains a path from x to y in and . Otherwise, by the previous paragraph, for some G-conjugate cʹ of c. Moreover, c and cʹ stabilize one-dimensional subspaces X and Xʹ, respectively, of . Since when n = 3, we deduce from Theorem 4.5 that contains an element . Hence contains a path from x to y in and . Therefore, .
It remains to determine lower bounds for when . Here, [20, Theorem 1.1 and Remark 1.3] and Proposition 3.10 show that the intersection graph of G is connected with diameter at least 5, and so Lemma 2.3 yields .
Acknowledgements
The author is grateful to Colva Roney-Dougal and Peter Cameron for helpful discussions regarding the original thesis on which this work is based and to Colva and anonymous referees for their useful comments on this paper.
Funding
This work was supported by the University of St Andrews (St Leonard’s International Doctoral Fees Scholarship and School of Mathematics and Statistics PhD Funding Scholarship) and EPSRC grant number EP/W522422/1.
References
1.
A.
Abdollahi
,
S.
Akbari
and
H. R.
Maimani
,
Non-commuting graph of a group
,
J. Algebra
298
(
2006
),
468
–
492
.
2.
M.
Aschbacher
and
G. M.
Seitz
,
Involutions in Chevalley groups over fields of even order
,
Nagoya Math. J.
63
(
1976
),
1
–
91
.
3.
W.
Bosma
,
J.
Cannon
and
C.
Playoust
,
The Magma algebra system. I. The user language
,
J. Symbolic Comput.
24
(
1997
),
235
–
265
.
4.
T.
Breuer
,
The GAP Character Table Library, Version 1.3.9
(
2024
).
5.
T.
Breuer
,
R. M.
Guralnick
and
W. M.
Kantor
,
Probabilistic generation of finite simple groups. II
,
J. Algebra
320
(
2008
),
443
–
494
.
6.
T.
Burness
,
R.
Guralnick
and
S.
Harper
,
The spread of a finite group
,
Ann. of Math. (2)
193
(
2021
),
619
–
687
.
7.
T. C.
Burness
,
R. M.
Guralnick
and
J.
Saxl
,
On base sizes for algebraic groups
,
J. Eur. Math. Soc. (JEMS)
19
(
2017
),
2269
–
2341
.
8.
T. C.
Burness
,
A.
Lucchini
and
D.
Nemmi
,
On the soluble graph of a finite group
,
J. Combin. Theory Ser. A
194
(
2023
), 105708.
9.
P. J.
Cameron
,
Graphs defined on groups
,
Int. J. Group Theory
11
(
2022
),
53
–
107
.
10.
P. J.
Cameron
,
S. D.
Freedman
and
C. M.
Roney-Dougal
,
The non-commuting, non-generating graph of a nilpotent group
,
Electron. J. Combin.
28
(
2021
), 15, .
11.
L.
Carlitz
,
Note on a quartic congruence
,
Amer. Math. Monthly
63
(
1956
),
569
–
571
.
12.
J. H.
Conway
,
R. T.
Curtis
,
S. P.
Norton
,
R. A.
Parker
and
R. A.
Wilson
,
Atlas of Finite groups.
Oxford University Press
,
Eynsham
,
1985
.
13.
E.
Crestani
and
A.
Lucchini
,
The non-isolated vertices in the generating graph of a direct powers of simple groups
,
J. Algebraic Combin.
37
(
2013
),
249
–
263
.
14.
B.
Csákány
and
G.
Pollák
,
The graph of subgroups of a finite group
,
Czechoslovak Math. J.
19
(
1969
),
241
–
247
.
15.
W.
de Launey
and
D.
Flannery
,
Algebraic Design Theory
,
Amer. Math. Soc
,
Providence, RI
,
2011
.
16.
L. E.
Dickson
,
Linear groups: With an Exposition of the Galois Field theory.
Dover Publications Inc
,
New York, NY
,
1958
.
17.
H.
Dietrich
,
M.
Lee
,
A.
Pisani
and
T.
Popiel
,
Explicit construction of the maximal subgroups of the Monster
, .
18.
H.
Dietrich
,
M.
Lee
and
T.
Popiel
.
The maximal subgroups of the Monster
, ,
2023
, .
19.
S. D.
Freedman
.
Diameters of graphs related to groups and base sizes of primitive groups - GAP and Magma code (thesis data).
August
2022
, doi:
.
20.
S. D.
Freedman
,
The intersection graph of a finite simple group has diameter at most 5
,
Arch. Math.
117
(
2021
),
1
–
7
.
21.
S. D.
Freedman
,
Diameters of graphs related to groups and base sizes of primitive groups.
,
University of St Andrews
,
May
2022
.
22.
S. D.
Freedman
,
The non-commuting, non-generating graph of a non-simple group
,
Algebr. Comb.
6
(
2023
),
1395
–
1418
.
23.
T.
Fritzsche
,
The depth of subgroups of
,
J. Algebra
349
(
2012
),
217
–
233
.
24.The GAP Group, GAP – Groups, Algorithms and Programming, Version 4.12.2
(
2022
).
25.
G.
Glauberman
,
Central elements in core-free groups
,
J. Algebra
4
(
1966
),
403
–
420
.
26.
D.
Gorenstein
,
R.
Lyons
and
R.
Solomon
,
The Classification of The Finite Simple groups. Number 3. Part I. Chapter A.
Amer. Math. Soc
,
Providence, RI
,
1998
.
27.
R. L.
Griess Jr.
,
Finite groups whose involutions lie in the center
,
Quart. J. Math. Oxford Ser. (2)
29
(
1978
),
241
–
247
.
28.
G.
Heide
,
J.
Saxl
,
P. H.
Tiep
and
A. E.
Zalesski
,
Conjugacy action, induced representations and the Steinberg square for simple groups of Lie type
,
Proc. Lond. Math. Soc. (3)
106
(
2013
),
908
–
930
.
29.
I. N.
Herstein
,
A remark on finite groups
,
Proc. Amer. Math. Soc.
9
(
1958
),
255
–
257
.
30.
M. D.
Hestenes
,
Singer groups
,
Canadian J. Math.
22
(
1970
),
492
–
513
.
31.
P. E.
Holmes
and
R. A.
Wilson
,
is a subgroup of the Monster
,
J. London Math. Soc. (2)
69
(
2004
),
141
–
152
.
32.
P. E.
Holmes
and
R. A.
Wilson
,
On subgroups of the Monster containing A5’s
,
J. Algebra
319
(
2008
),
2653
–
2667
.
33.
B.
Huppert
,
Endliche Gruppen. I.
Springer-Verlag
,
Berlin-New York, NY
,
1967
.
34.
R.
Lidl
and
H.
Niederreiter
,
Finite fields.
CUP
,
Cambridge
,
1997
, 2nd Edn..
35.
M. W.
Liebeck
and
J.
Saxl
,
On point stabilizers in primitive permutation groups
,
Comm. Algebra
19
(
1991
),
2777
–
2786
.
36.
M. W.
Liebeck
,
J.
Saxl
and
G. M.
Seitz
,
Subgroups of maximal rank in finite exceptional groups of Lie type
,
Proc. London Math. Soc. (3)
65
(
1992
),
297
–
325
.
37.
M. W.
Liebeck
and
A.
Shalev
,
Simple groups, permutation groups and probability
,
J. Amer. Math. Soc.
12
(
1999
),
497
–
520
.
38.
A.
Lucchini
and
D.
Nemmi
,
On the connectivity of the non-generating graph
,
Arch. Math.
118
(
2022
),
563
–
576
.
39.
S.
Perlis
,
Theory of matrices
,
Dover Publications Inc
,
New York, NY
,
1991
.
40.
J. S.
Rose
,
On finite insoluble groups with nilpotent maximal subgroups
,
J. Algebra
48
(
1977
),
182
–
196
.
41.
M.
Suzuki
,
On a class of doubly transitive groups
,
Ann. of Math. (2)
75
(
1962
),
105
–
145
.
42.
R. A.
Wilson
,
The Finite Simple Groups
,
Springer-Verlag London, Ltd
,
London
,
2009
.
43.
R. A.
Wilson
. Maximal subgroups of sporadic groups,
Finite Simple Groups: Thirty Years of the Atlas and Beyond, Amer. Math.
Soc
.,
Providence, RI
,
2017
,
57
–
72
.
© The Author(s) 2025. Published by Oxford University Press.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (
https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.